Open access peer-reviewed chapter

Twenty-First Century Glucocorticoid Receptor Molecular Biology

Written By

Cheng Wang, Roel Oldenkamp, Ronald J.W. Oellers and Colin Logie

Submitted: 01 June 2017 Reviewed: 27 October 2017 Published: 20 December 2017

DOI: 10.5772/intechopen.72016

From the Edited Volume

Corticosteroids

Edited by Ali Gamal Al-kaf

Chapter metrics overview

1,352 Chapter Downloads

View Full Metrics

Abstract

Glucocorticoids are central to homeostasis as a function of the circadian cycle, temporally preceding circulating adrenaline concentration circadian fluctuations. Virtually, all cell types express the glucocorticoid receptor (GR). GR is a transcription factor that activates gene expression by binding to enhancers. Intriguingly, not all cell types respond to GR stimulation in the same fashion at the molecular level. This indicates that GR activity is subject to epigenetic control. We discuss the molecular basis for epigenetic control of GR action at the genomic level, including the concept of topologically associating domains which may restrain the roaming range of distal enhancers. Furthermore, much evidence indicates that GR can repress gene expression programs. We therefore discuss current concepts of the molecular basis of GR-mediated gene expression repression, including non-genomic mechanisms that involve mRNA destabilization.

Keywords

  • glucocorticoid receptor
  • glucocorticoid response element
  • chromosome conformation
  • epigenetics
  • non-genomic action
  • RNA decay

1. Introduction

Glucocorticoids (GCs) are steroids derived from cholesterol that are mainly produced in the adrenal cortex, under the control of the hypothalamic–pituitary–adrenal axis. Due to their lipophilic nature, GCs can traverse cellular membranes and thus enter any cell. Physiologically, GCs show circadian oscillations in man, peaking at 06:00 before we wake up and then dropping until 00:00, when their levels start to rise again. Adrenaline, a catecholamine that is produced by the adrenal medulla, follows this trend with a lag of about 2 hours [1]. Ontogenetically, GC levels increase during the final weeks of human gestation and in the post-natal period. This not only stimulates gluconeogenesis, but also perinatal lung maturation [2] and many other physiological processes [3, 4, 5]. Furthermore, GCs are part of an emotion (stress, fear, and arousal) processing pathway in the brain that impacts memory and aspects of behavior that are controlled by the central nervous system [6, 7, 8].

Importantly, from a medical point of view, GCs and their synthetic analogues have strong immunosuppressive properties. Because of this, synthetic GCs belong to the top 50 World Health Organization essential medicines. Prednisone, dexamethasone, and triamcinolone are used to treat a wide range of (auto)inflammatory conditions as well as hematopoietic malignancies. The anti-inflammatory effect of GCs is due to regulation of cell survival and immune signaling molecules such as chemokines, interleukins, and cytokines such as TNFα [9, 10]. GCs are often well accepted as a long-term treatment, making them irreplaceable for medical use. Nevertheless, synthetic glucocorticoid (over)use has a number of side effects that usually involve homeostasis and tissue maintenance [11, 12]. To mitigate such side effects, a detailed understanding of the molecular mode of action of GCs is a necessity. Hence, understanding the molecular mechanisms through which GCs exert their biological function has been a highly active research field in the past century.

The glucocorticoid receptor (NR3C1, abbreviated here as GR) is a sequence-specific DNA-binding transcription factor that is expressed in virtually every human cell type. Hence, almost every tissue is potentially responsive to GCs through gene expression modulation. Since the molecular responses to GCs of given tissues are different, it is thought that epigenetic programming during cellular differentiation underlies the cell-specific GC responses [13]. Below, we will review recent developments in epigenetic research relevant to cell-specific GC response mechanisms. In the last section of this chapter, we will review recent research results that support the notion that non-genomic effects of GCs may be very important too.

Advertisement

2. Chromosome architecture and epigenetic control of glucocorticoid responses: DNA accessibility

Eukaryotic transcription factors (TFs) bind to regulatory DNA elements commonly called “cis-acting elements” to modulate the transcription rates of their target genes. Cis-acting elements can be located at (i) gene promoters, where mRNA transcription starts, or (ii) at enhancers, which can be located hundreds of thousands of nucleotides away from their target gene promoters, or (iii) at boundary elements that flank chromosome domains and function to restrict enhancer activity within individual topologically associated chromosome domains [13].

In order to determine the locations where TFs bind on chromosomes, a technique called chromatin immunoprecipitation (ChIP) was developed in the 1990s. ChIP is based on formaldehyde crosslinking of TFs to DNA, followed by DNA co-immunoprecipitations using antibodies directed against the TF protein [14]. Initially, PCR was used to analyze the co-immunoprecipitated DNA, using the enrichment of putative TF target sites relative to “control” chromosomal regions. Nowadays, the co-immunoprecipitated DNA fragments are prepared as DNA libraries that can be sequenced on next-generation sequencing (NGS) platforms, followed by computational mapping of the obtained reads to a reference genome [15].

Currently, more than 20 human and mouse genome-wide GR occupation profiles are available. These reveal a high degree of GR-binding variability [16]. Grøntved et al. showed that a majority (83%) of GR-DNA binding sites in mouse liver cells are liver cell-specific, while only 0.5% of events are shared between all analyzed cell-types [17]. This suggests that there is a complex and dynamic epigenetic component to GR binding that underlies the differences in GR-mediated transcriptional regulation across cell types.

The first level of epigenetic regulation is rather well defined by DNA being wrapped, or not, around histones to form nucleosomes every ~190 bp [18, 19]. Low nucleosome occupancy can be measured as DNaseI hypersensitivity, because accessible free DNA is more prone to DNaseI endonuclease cleavage than DNA wrapped around nucleosomes [20, 21].

DNA accessibility is an important indicator for GR binding. Early studies indicated that GR binding increases DNA accessibility to DNaseI [22, 23] and it was therefore concluded that GR “opens up” chromatin. Although this is true, more recent research indicates that the a majority of chromosomal GR-binding sites coincide with pre-existing hypersensitive DNA stretches, whose DNaseI accessibility profile is further modulated by GR activity, as first reported on a genome-wide level by John et al. [24, 25, 26, 27] (Figure 1). Grøntved et al. indicated that 62% of glucocorticoid receptor-binding sites are occupied by the transcription factor C/EBP in mouse liver tissue and that C/EBP maintains chromatin accessibility before GC treatment [17]. Furthermore, it was shown that in HeLa cells, 88% of GR-binding sites are already occupied by the lysine acetyltransferase p300 transcription co-activator prior to GC treatment [27, 28]. Altogether, the available evidence indicates that GR-mediated transcriptional control is dependent on other TFs that establish baseline chromatin accessibility profiles in a cell-type specific manner, as exemplified by FoxA1 [29]. However, one single pioneer TF is unlikely to be the sole key to differential use of GR response elements by different cell lineages, or by the same cell type under different conditions. Rather, combinations of DNA sequence-specific transcription factors may act together as “reciprocal pioneers” in an environmentally cued fashion [30, 31, 32, 33].

Figure 1.

GR binds to GREs at several DNaseI hypersensitive locations within the TSC22D3/GILZ locus on human chromosome X. This can increase p300 histone acetyltransferase occupancy, H3K27ac marking, and DNaseI hypersensitivity. Notably, occupancy by RNA polymerase II is dramatically increased upon 4 hours of GC treatment, indicating transcription activation. Histone H3 lysine modifications are indicated (H3K27 acetylation, H3K4 mono-methylation, H3K4 tri-methylation). Data are from HeLa cells, Rao et al. [27].

In summary, in a given cell type, GR generally binds to a predetermined set of nucleosome free regions within enhancers that are marked by lineage determining TFs, and GR only rarely binds at sites with very low initial levels of DNA accessibility (Figure 1) [28, 34]. Intriguingly, GR appears to associate for rather short times with its cognate sites in vivo, with reported DNA residence times in the order of seconds [35, 36, 37, 38, 39]. GR binding usually results in increased histone acetylation [27] (Figure 1).

Advertisement

3. Chromosome architecture and epigenetic control of glucocorticoid responses: topologically associated domains

Over the last decade, a lot of effort was invested in mapping active cis-regulatory enhancer elements to susceptible promoters. This is especially relevant in GR-mediated transcriptional regulation, because the majority of GR-bound cis-acting DNA elements are enhancers that are located many kilobases away from the promoters of GR-responsive genes [17, 24]. An important contributor in the identification of enhancer-promoter interactions was the development of nuclear proximity-based chromosome conformation capture (3C) technology in 2002 [40]. In brief, interacting DNA regions are fixed by formaldehyde through DNA-protein-DNA crosslinks. The crosslinked chromatin is then digested using restriction enzymes and the digested ends are ligated to obtain DNA circles that harbor sequences from interacting DNA regions. In the original 3C protocol, which is considered a “one-to-one” approach, interactions between two defined genomic loci are assessed by quantitative polymerase chain reaction (RT-qPCR) using locus-specific primers. Circularized Chromatin Conformation Capture (4C), is a “one-to-all” approach that implements a second round of restriction and ligation to obtain small DNA circles which are suitable for inverse PCR amplification to identify the genome-wide DNA interactions of one defined viewpoint locus with any other chromosomal loci [41, 42]. The most recent technical development in 3C technologies is the establishment of chromosome capture followed by high-throughput sequencing (Hi-C) [43]. Crosslinked DNA is digested, labeled with biotin, and re-ligated resulting in a biotin-labeled 3C library. Ligated circles are sheared, purified, and subsequently analyzed using NGS. Hi-C is an “all-to-all” approach because it potentially identifies all possible genome-wide DNA interactions. Capture Hi-C is a further modified version of Hi-C that uses immobilized custom DNA probes and DNA hybridization to enrich for specific loci interactions present in a Hi-C library [44].

A fascinating feature of nuclear chromosome organization is its hierarchical character, containing several layers of compartmentalization. Analyses of Hi-C contact matrices confirm the existence of a first level of organization, namely the occurrence of chromosome territories [45] that were previously described in microscopy-based studies [46]. At the next level, individual chromosomes are partitioned into multi-megabase “A” and “B” compartments that have a propensity to cluster separately. “A” compartments tend to display a euchromatin profile, being gene-rich, transcriptionally active, and accessible. “B“ compartments are generally gene-poor with a tendency to be more heterochromatic, transcriptionally inactive, and less accessible. Hi-C maps with improved resolution, mainly obtained through increased sequencing depth and the use of different restriction enzymes, reveal the partitioning of A and B compartments into so-called sub-Mb-sized topologically associated domains (TADs) [47]. TADs are defined by their tendency to favor internal rather than external DNA interactions. Hence, the TAD hypothesis states that TADs are flanked by left and right boundaries and that enhancers mainly interact with promoters and enhancers within their TAD, but not outside of it. It is currently thought that TADs consist of dynamic sub-Mb chromatin fiber loops that undergo continuous remodeling, among others through RNA polymerase II passage.

TADs are highly conserved between different cell lineages [48], indicating that TADs may be universal functional chromosomal units that serve as a platform within which cis-regulatory elements are spatially brought together with their susceptible promoter element. The basis of TAD loops is highly enriched for CCCTCF-binding factor (CTCF) [47] (Figure 2). CTCF is known as a transcriptional regulator that functionally segregates chromosomal TADs by inhibiting enhancer-promoter interactions [49]. Importantly, the majority of mammalian TAD loops are flanked by a pair of convergent CTCF motifs that mark the TAD’s left and right boundaries [50]. Deletion or inversion of CTCF sites can alter TAD architecture and therefore result in dysregulated enhancer-promoter interactions [51]. Moreover, CTCF depletion disrupts TAD boundaries [52] and impacts gene expression [53]. Dysregulation of CTCF is associated with improper gene regulation during development and oncogenesis [54, 55].

Figure 2.

A model depicting long-range transactivation after glucocorticoid stimulation. (A) Linear overview of cis-acting element organization. Convergent CTCF motifs define TAD boundaries that restrict promoter-enhancer interactions. A schematic contact matrix of a virtual Hi-C experiment is shown as an interaction heatmap. (B) GR induces transcription through binding of a pre-configured locus without affecting its spatial chromosome architecture. Low and high levels of enhancer H3K27 acetylation are depicted by light and dark rectangles, respectively. TAD: topologically associating domain, GR: glucocorticoid receptor, and TSS: transcription start site. See also Ref. [13].

The cohesin complex co-localizes with CTCF when assayed by ChIP [56, 57]. Cohesin rings are composed of the core subunits SMC1, SMC3, RAD21, and STAG [58, 59]. Cohesin is most likely loaded onto its chromatid substrate by the NIPBL2/Mau2 cohesin-loading complex, which is enriched at transcription start sites (TSS) [60, 61]. Conversely, cohesin release from chromatids is facilitated by WAPL [62]. Depletion of cohesin leads to altered short-range chromatin interactions, while global TAD organization seemingly persists, suggesting that cohesin and CTCF play different mechanistic roles in TAD formation [63]. Indeed, while inhibiting cohesin loading (by inhibiting cohesin loading factors) inhibits the formation of topologically associated domains, inhibiting cohesin release by inhibiting WAPL restricts loop extension [64]. In the absence of both CTCF and WAPL, cohesin accumulates in up to 70 kilobase-long regions at the 3′-ends of active genes, in particular, if these converge on each other [60, 61]. Cohesin can be moved along chromosomes through RNA polymerase II translocation along its template in yeast and human; this indicates evolutionary conservation of the translocation of Cohesin rings during RNA polymerase II passage.

A quantitative model of “chromatin loop extrusion” was proposed that explains the dynamic features of TADs rather well [50, 65, 66]. Very recently, looping was studied in the monocytic leukemia cell line THP1 that can differentiate into macrophage-like cells. About 16,000 chromatin loops were detected in both cell types and, using stringent selection criteria, 217 were found to be “dynamic” [67]. This indicates that although loss and gain of TAD loops can occur naturally as cells adapt their gene expression landscape, it is not an obligate step in gene activation/repression. Indeed, Hi-C results obtained in parallel in eight primary human hematopoietic cell types show high correspondence [68].

In 2009, long-range interactions involving GR-bound cis-acting DNA sequences were identified in mouse cells using a modified 3C technique [70]. An interaction that spans 30 kb was detected between a GR-binding site in the Lcn2 gene and the promoter of the Ciz1 gene. This interaction may be responsible for GC-mediated Lcn2 and Ciz1 transcription induction in mouse mammary epithelial adenocarcinoma 3134 cells [69]. In 2011, the same research group reported that “the predominant hormone-induced changes for Lcn2-contacting loci can be attributed to an increased frequency of pre-existing interactions” [70]. More recently, the 4C approach and genome-wide chromatin structure analysis were applied to characterize GR-associated DNA interactions [71]. In the 3134 murine cell line, this showed that activated GR response elements can interact with a downstream enhancer of the Tsc22d3 transcription repressor gene, whose transcription is strongly upregulated by glucocorticoids. See also Figure 1 where human TSC22D3 is shown. However, upon glucocorticoid receptor activation, contact intensities changed two-fold at most [71].

Theoretically, there are two types of models for transcription factor (TF)-mediated gene regulation at the level of chromatin organization and chromosome folding. In the first type, repressed loci reside in a silent and inaccessible chromatin state with a low enhancer-promoter interaction frequency. Binding of TFs to distal cis-regulatory elements would then enhance the accessibility of the locus for other TFs to bind the enhancers and promoters, and consequently, increased interaction between promoter and enhancer elements would alter gene expression [72, 73]. In the second type of models, the locus is dynamically pre-configured in 3D through boundary-boundary interactions controlled by CTCF and cohesin dynamics that insure that TFs can rapidly exert stimulatory or repressive effects on transcription [74] (Figure 2). In this model, TFs hardly affect enhancer-promoter interaction frequencies, although they do affect the histone-borne epigenetic marks such as H3K27 acetylation (see Figure 1). Currently, available data suggest that GR-responsive loci fit the second type of models, since dexamethasone-mediated GR activation does not greatly alter TAD structure [70, 71] (Figure 2).

Advertisement

4. GR-binding site sequences and GR-mediated transrepression

The oligomerization state and quaternary structure of GR protein on DNA is thought to influence the activity of cis-acting GR-binding DNA elements. Experimentally determined glucocorticoid receptor DNA-binding sites have been broadly classified as “simple,” “composite” or “tethering.” In the “simple” case, homodimers of GR trans-activate genes by binding to canonical GR response elements (GREs) and consequently recruit transcription co-activators [75]. In the composite DNA motif case, repression and activation are both possible outcomes. Finally, “tethering” is a DNA-binding mode whereby GR does not directly bind specific DNA sequences; instead, it is indirectly tethered to DNA by another TF via protein-protein interactions. Tethering was historically proposed to be the main mechanism of GC-induced GR-mediated transcription repression.

Canonical GREs, mineralocorticoid, progesterone, and testosterone receptor-binding sites are virtually identical, being composed of two inverted pseudo-palindromic repeats separated by a spacer sequence of three bases (GRACANNNTGTYC) [76, 112]. Spacer sequence length has been proposed to be important to maintain GR’s dimerization state [77, 78]. Furthermore, it has been suggested that allosteric DNA plasticity in the GR recognition sequences influences the conformational state of GR and, thereby, its spatiotemporal regulatory character [79, 80]. However, Presman et al., shone new light on this paradigm as real-time imaging suggests that GR tetramerizes at GREs [81]. Furthermore, in another key publication, Presman et al. used GR point mutations to confirm that trans-repression and transactivation by GR are two functions that can be separated genetically, whereby loss of transactivation potential though impaired homodimerization did not always co-occur with loss of trans-repression potential [82].

The application of single-base resolution TF ChIP technology, attained by inclusion of a lambda exonuclease digestion step in the ChIP protocol (ChIP-exo), was used to reveal that many GR-bound half-sites (GRACA) coincide with recognition sequences of unrelated TFs at composite elements [83, 84]. For instance, Lim et al. revealed co-localization with liver-specific TF-binding sites, explaining part of GR’s liver cell-specific binding profiles [84]. The molecular mode of regulation at composite sites still remains to be elucidated, although it was hypothesized to fit a model in which only the co-association of the involved TFs results in productive DNA binding, as seen for classical heterodimeric TFs [85, 86].

Next to its ability to bind half-sites, monomeric GR has been reported to counter the effects of other TFs through protein-protein tethering which would result in trans-repression [87, 88]. One such proposed GR-tethering partner is the activator protein 1 (AP-1) heterodimer made-up of heterodimers of bZIP TF family members. A second major proposed GR tethering partner is NF-κB, a TF that consists of heterodimers of RELA and RELB with NFKB1 and NFKB2 subunits [89, 90, 91, 92].

For long, AP-1 and NF-κB tethering of GR to DNA were considered the dominant mechanism for GR-mediated trans-repression of transcription, through “on-DNA” repression of the GR tethering TF’s transcription activation potential, as reviewed by Glass and Saijo [93]. Genomic studies showed a significant reduction of GR association upon AP-1 loss, but a majority of regulatory scenarios could neither be disentangled nor rationalized through genome-wide ChIP analyses [94]. Indeed, recent experiments indicate that the mode of GR “trans-repression” is still not fully understood. For instance, Oh et al., showed that activation of GR after LPS treatment caused similar gene repression as activation of GR before LPS treatment, and that DNA occupancy by GR was not predictive of gene expression repression, contradicting the “trans-repression by tethering” model. Rather, GR activation was found to directly induce the expression of inhibitors of NF-κB (and AP-1) and this was proposed to cause genome-wide blockade of NF-κB interaction with chromatin [95]. This suggests that protein tethering leading to DNA-bound monomeric GR trans-repression can only account for a minority of repressive events [96]. Indeed, single-molecule imaging suggests that tethering can account for only ~3% of DNA recruitment events [35].

In yet another twist of the GR tethering saga, Weikum et al. showed that GR associates with a GRE half-site that is located within an AP-1 recognition element, even in the absence of AP-1 [97]. Since AP-1 occupancy was not directly required for GR-mediated trans-repression, Weikum et al. proposed that AP-1 establishes an accessible chromatin state for subsequent GR binding to the half-sites which results in transcription repression [34]. Whether AP-1 trans-repression by GCs relies on co-repressor recruitment [98, 92] or rather on exclusion of other TFs and their co-activators is an unresolved issue at this point in time.

Advertisement

5. Non-genomic mechanisms of gene regulation by glucocorticoids

The classical model for GR action involves ligand-dependent release from a repressive HSP90 complex followed by genomic DNA binding and consequent transcription modulation [12, 75, 99101]. However, over the years, non-genomic physiologically relevant GR responses have been proposed, as reviewed by Boldizsar et al. [102]. These include direct membrane binding effects of (synthetic) glucocorticoids, a putative non-GR membrane-associated receptor [103], functional interactions of GR with proteins involved in signal transduction such as kinases and phosphatases [104, 105], and mitochondrial GR translocation as a mechanism leading to T-cell apoptosis [106]. An advantage of non-genomic regulation over genomic regulation of gene expression is that non-genomic regulation can take place much faster than the transcription-translation process, which often take >20 minutes to begin to change a cell’s molecular composition [107, 108, 109].

Over the past decades, it has become apparent that non-genomic mechanisms may also play vital roles in GC action, particularly in the context of immune cell regulation [108]. The mechanism we will review below concerns the apparent capacity of GR to bind to RNA.

There are reports that the growth arrest-specific 5 transcript (GAS5), which is a non-coding RNA, can sequester GR, as well as progesterone and androgen receptors, away from their genomic sites of action by acting as a “GRE decoy” [110, 111, 112]. This GAS5-dependent GC inhibitory pathway appears to also be active in some immune cells [113, 114]. Although there are no crystal structures of GR bound to RNA, such structures have been modeled [110].

On the other hand, evidence was published that GCs affect the turnover of specific mRNAs. Regulation of mRNA stability is an intricate process controlled by a complex set of interaction between phosphorylation-mediated signaling pathways like the phosphorylation of UPF1 or SMG-2, which together with cis-regulatory RNA elements accelerate an mRNA’s decay rate [107, 115, 116, 117, 118]. RNA cis-acting elements that regulate mRNA stability are usually found in their 5′ and 3′ untranslated regions (UTRs) [119, 120]. The most widely found sequences in the 3′ UTRs of unstable mRNAs belong to the adenylate-uridylate-rich elements, consisting of AUUUA ribonucleotide sequences [119]. It has been proposed that GCs can accelerate mRNA decay by inducing the transcription of genes that code for protein factors implicated in mRNA decay. One such example being the gene that codes for tristetraprolin (TTP, also known as ZFP36), which is inducible by GCs under some circumstances [121, 122]. Pro-inflammatory factor mRNAs indeed display differential half-lives through such an indirect GC-induced mechanism, an example of which is TNFα [122, 123].

Strikingly, in addition to upregulating the expression of mRNA decay factors, it would appear that GR can act directly as a ligand-dependent activator of mRNA decay. In 1999, a 5′ UTR RNA element was reported to be of particular importance for GC regulation of the expression of the MCP-1/CCL2 inflammatory chemokine [124]. In 2007, it was first reported that GR binds specifically to CCL2 mRNA, to cause its decay [125]. In 2011, an RNA immunoprecipitation protocol was employed to define an RNA motif that recruits GR and the 5′ UTRs of CCL2 and CCL7 mRNAs [126]. The mechanism of GR binding to an mRNA to mediate its decay was termed “GR-mediated mRNA decay” (GMD) by Park et al. in 2015 [127]. This research group investigated how GMD occurs. They reported that GMD is a distinct mRNA decay pathway that shares factors with other forms of RNA decay [128, 129]. GMD depends on a number of proteins that have to be recruited to the mRNA. These include GR, PNRC2, UPF1, DCP1A, HRSP12, and YBX1 which then instigate rapid mRNA degradation (Figure 3). PNRC2 and UPF1 are known to bind to each other to bring RNA helicase activity into the complex. Another pair of factors that are known for their ability to degrade mRNA are DCP1A, which promotes mRNA decapping by DCP1 activity, and HRSP12, an endoribonuclease that can attack mRNA [129]. Although exciting, GMD still needs to be confirmed by unbiased approaches such as genome-wide transcriptomic comparisons of nascent RNA and steady-state RNA which have the capacity to simultaneously report mRNA transcription and decay rates [130].

Figure 3.

Model of the assembly and composition of the glucocorticoid mediated mRNA decay pathway as described by Park et al. [129]. GR with bound GC recruits PNRC2 and DCP1A together with UPF1 to the 5′ UTR of a target mRNA to form Complex I. HRSP12 and YBX1 are then recruited to form Complex II and mRNA decay is performed by exonucleases.

Advertisement

6. Conclusion

Over the past decade, GR action has been studied at the molecular level in model systems using DNA accessibility assays, GR ChIP, epigenetic profiling of histone-borne epigenetic marks, transcriptome profiling, and RNA immunoprecipitation. Furthermore, chromosome conformation capture assays have been deployed to investigate the impact of GC signaling on chromosome domain topology.

In the cases where it was studied, GR was found to bind for less than a minute to its genomic targets. GR does not appear to affect the configuration of the topologically associated domains to which it binds. It therefore appears that GR binds to loci where enhancers and promoters are dynamically pre-configured in three-dimensional space. The observation that GR complies with chromosome conformation rather than influencing it offers the exciting perspective of being able to map intergenic GRE’s, which are often located very far away from their target promoters, to TADs. The genes encompassed by these TADs can then be earmarked as potential GR target genes, a hypothesis that can be confirmed by monitoring their expression upon GC exposure.

Results obtained by many laboratories suggest that GR is dependent on other pioneer transcription factors to access its response elements in chromosomal DNA. Co-pioneer factor combinations appear to be cell-type-specific lineage determining TFs, largely explaining the tissue-specific responses elicited by GCs. Furthermore, much evidence indicates that GR is not only restricted to the classical inverted repeat steroid response element, but can also bind to a variety of DNA sequences that only encompass one half site. Furthermore, the concept that GR is tethered indirectly to DNA via other TFs, whose activity it would then repress “on DNA,” is no longer the only model to explain trans-repression in the field. Indeed, other genomic and non-genomic interactions may explain the repression of NFKB and AP-1 target genes observed upon GC exposure.

Interestingly, GR itself appears to be subject to miRNA-mediated regulation, as recently reviewed [131].

Excitingly, following on early reports of RNA binding, it was reported multiple times that GR is also an mRNA-binding protein that induces mRNA decay. A particular target for this pathway are CCL chemokine family mRNAs that have long been known to undergo a dramatic down-regulation upon GC exposure.

Altogether, we conclude that although much effort has been invested in glucocorticoid research since the discovery in the 1940s that glucocorticoids are anti-inflammatory wonder drugs, much remains to be discovered about the molecular mechanisms of action of glucocorticoids.

Advertisement

Abbreviations

3CChromosome Conformation Capture
4CCircularized Chromatin Conformation Capture
ChIPChromatin immunoprecipitation
DHSDNaseI hypersensitive site
GCglucocorticoid
GMDGR-mediated mRNA decay
GRGlucocorticoid receptor
GREGR response elements
Hi-CChromosome conformation capture with high-throughput sequencing
LPSLipopolysaccharides
NGSNext-generation sequencing
TADTopologically associating domain
TFTranscription factor
TNFαTumor necrosis factor
TSSTranscription start site
UTRmRNA untranslated region

References

  1. 1. Scheer FA et al. Impact of the human circadian system, exercise, and their interaction on cardiovascular function. Proceedings of the National Academy of Sciences of the United States of America. 2010;107(47):20541-20546
  2. 2. Roberts D et al. Antenatal corticosteroids for accelerating fetal lung maturation for women at risk of preterm birth. Cochrane Database of Systematic Reviews. 2017;3:CD004454
  3. 3. Liggins GC. The role of cortisol in preparing the fetus for birth. Reproduction, Fertility, and Development. 1994;6(2):141-150
  4. 4. Gillies GE et al. Enduring, sexually dimorphic impact of in utero exposure to elevated levels of glucocorticoids on midbrain dopaminergic populations. Brain Sciences. 2016;7(1)
  5. 5. Alcantara-Alonso V et al. Corticotropin-releasing hormone as the homeostatic rheostat of Feto-maternal Symbiosis and developmental programming in utero and neonatal life. Frontiers in Neuroendocrinology (Lausanne). 2017;8:161
  6. 6. Varney NR, Alexander B, MacIndoe JH. Reversible steroid dementia in patients without steroid psychosis. The American Journal of Psychiatry. 1984;141(3):369-372
  7. 7. Cahill L, McGaugh JL. Mechanisms of emotional arousal and lasting declarative memory. Trends in Neurosciences. 1998;21(7):294-299
  8. 8. Hunsberger JG et al. Cellular mechanisms underlying affective resiliency: The role of glucocorticoid receptor- and mitochondrially-mediated plasticity. Brain Research. 2009;1293:76-84
  9. 9. Savino W et al. Hormonal control of T-cell development in health and disease. Nature Reviews. Endocrinology. 2016;12(2):77-89
  10. 10. Zen M et al. The kaleidoscope of glucorticoid effects on immune system. Autoimmunity Reviews. 2011;10(6):305-310
  11. 11. Allen DB. Effects of inhaled steroids on growth, bone metabolism, and adrenal function. Advances in Pediatrics. 2006;53:101-110
  12. 12. Cain DW, Cidlowski JA. Immune regulation by glucocorticoids. Nature Reviews. Immunology. 2017;17(4):233-247
  13. 13. Logie C, Stunnenberg HG. Epigenetic memory: A macrophage perspective. Seminars in Immunology. 2016;28(4):359-367
  14. 14. Orlando V. Mapping chromosomal proteins in vivo by formaldehyde-crosslinked-chromatin immunoprecipitation. Trends in Biochemical Sciences. 2000;25(3):99-104
  15. 15. Thomas R et al. Features that define the best ChIP-seq peak calling algorithms. Briefings in Bioinformatics. 2017;18(3):441-450
  16. 16. Sacta MA, Chinenov Y, Rogatsky I. Glucocorticoid signaling: An update from a genomic perspective. Annual Review of Physiology. 2016;78:155-180
  17. 17. Grøntved L et al. C/EBP maintains chromatin accessibility in liver and facilitates glucocorticoid receptor recruitment to steroid response elements. The EMBO Journal, 2013. 32(11): p. 1568-1583
  18. 18. Valouev A et al. Determinants of nucleosome organization in primary human cells. Nature. 2011;474(7352):516-520
  19. 19. Jiang C, Pugh BF. Nucleosome positioning and gene regulation: Advances through genomics. Nature Reviews. Genetics. 2009;10(3):161-172
  20. 20. Gross DS, Garrard WT. Nuclease hypersensitive sites in chromatin. Annual Review of Biochemistry. 1988;57:159-197
  21. 21. Dorschner MO et al. High-throughput localization of functional elements by quantitative chromatin profiling. Nature Methods. 2004;1(3):219-225
  22. 22. Richard-Foy H, Hager GL. Sequence-specific positioning of nucleosomes over the steroid-inducible MMTV promoter. The EMBO Journal. 1987;6(8):2321-2328
  23. 23. Reik A, Schutz G, Stewart AF. Glucocorticoids are required for establishment and maintenance of an alteration in chromatin structure: Induction leads to a reversible disruption of nucleosomes over an enhancer. The EMBO Journal. 1991;10(9):2569-2576
  24. 24. John S et al. Chromatin accessibility pre-determines glucocorticoid receptor binding patterns. Nature Genetics. 2011;43(3):264-268
  25. 25. Lorzadeh A et al. Nucleosome density ChIP-Seq identifies distinct chromatin modification signatures associated with MNase accessibility. Cell Reports. 2016;17(8):2112-2124
  26. 26. Reddy TE et al. The hypersensitive glucocorticoid response specifically regulates period 1and expression of circadian genes. Molecular and Cellular Biology. 2012;32(18):3756-3767
  27. 27. Rao NA et al. Coactivation of GR and NFKB alters the repertoire of their binding sites and target genes. Genome Research. 2011;21(9):1404-1416
  28. 28. Kuznetsova T et al. Glucocorticoid receptor and nuclear factor kappa-B affect three-dimensional chromatin organization. Genome Biology. 2015;16:264
  29. 29. Eeckhoute J et al. Cell-type selective chromatin remodeling defines the active subset of FOXA1-bound enhancers. Genome Research. 2009;19(3):372-380
  30. 30. Adam RC, Fuchs E. The yin and Yang of chromatin dynamics in stem cell fate selection. Trends in Genetics. 2016;32(2):89-100
  31. 31. Drouin J. 60 YEARS OF POMC: Transcriptional and epigenetic regulation of POMC gene expression. Journal of Molecular Endocrinology. 2016;56(4):T99-T112
  32. 32. Swinstead EE et al. Steroid receptors reprogram FoxA1 occupancy through dynamic chromatin transitions. Cell. 2016;165(3):593-605
  33. 33. Zaret KS, Lerner J, Iwafuchi-Doi M. Chromatin scanning by dynamic binding of Pioneer factors. Molecular Cell. 2016;62(5):665-667
  34. 34. Biddie SC et al. Transcription factor AP1 potentiates chromatin accessibility and glucocorticoid receptor binding. Molecular Cell. 2011;43(1):145-155
  35. 35. Gebhardt JC et al. Single-molecule imaging of transcription factor binding to DNA in live mammalian cells. Nature Methods. 2013;10(5):421-426
  36. 36. Stasevich TJ et al. Cross-validating FRAP and FCS to quantify the impact of photobleaching on in vivo binding estimates. Biophysical Journal. 2010;99(9):3093-3101
  37. 37. Mikuni S, Tamura M, Kinjo M. Analysis of intranuclear binding process of glucocorticoid receptor using fluorescence correlation spectroscopy. FEBS Letters. 2007;581(3):389-393
  38. 38. Morisaki T et al. Single-molecule analysis of transcription factor binding at transcription sites in live cells. Nature Communications. 2014;5:4456
  39. 39. Groeneweg FL et al. Quantitation of glucocorticoid receptor DNA-binding dynamics by single-molecule microscopy and FRAP. PLoS One. 2014;9(3):e90532
  40. 40. Dekker J et al. Capturing chromosome conformation. Science. 2002;295(5558):1306-1311
  41. 41. Simonis M et al. Nuclear organization of active and inactive chromatin domains uncovered by chromosome conformation capture-on-chip (4C). Nature Genetics. 2006;38(11):1348-1354
  42. 42. Olivares-Chauvet P et al. Capturing pairwise and multi-way chromosomal conformations using chromosomal walks. Nature. 2016;540(7632):296-300
  43. 43. van Berkum NL et al. Hi-C: A method to study the three-dimensional architecture of genomes. Journal of Visualized Experiments. 2010(39)
  44. 44. Hughes JR et al. Analysis of hundreds of cis-regulatory landscapes at high resolution in a single, high-throughput experiment. Nature Genetics. 2014;46(2):205-212
  45. 45. Lieberman-Aiden E et al. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science. 2009;326(5950):289-293
  46. 46. Bolzer A et al. Three-dimensional maps of all chromosomes in human male fibroblast nuclei and prometaphase rosettes. PLoS Biology. 2005;3(5):e157
  47. 47. Dixon JR et al. Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature. 2012;485(7398):376-380
  48. 48. Razin SV, Ulianov SV. Gene functioning and storage within a folded genome. Cellular & Molecular Biology Letters. 2017;22:18
  49. 49. Liu M et al. Genomic discovery of potent chromatin insulators for human gene therapy. Nature Biotechnology. 2015;33(2):198-203
  50. 50. Sanborn AL et al. Chromatin extrusion explains key features of loop and domain formation in wild-type and engineered genomes. Proceedings of the National Academy of Sciences of the United States of America. 2015;112(47):E6456-E6465
  51. 51. Guo Y et al. CRISPR inversion of CTCF sites alters genome topology and enhancer/promoter function. Cell. 2015;162(4):900-910
  52. 52. Nora EP et al. Targeted degradation of CTCF decouples local insulation of chromosome domains from genomic compartmentalization. Cell. 2017;169(5):930-944 e22
  53. 53. Nikolic T et al. The DNA-binding factor Ctcf critically controls gene expression in macrophages. Cellular & Molecular Immunology. 2014;11(1):58-70
  54. 54. Lettice LA et al. Enhancer-adoption as a mechanism of human developmental disease. Human Mutation. 2011;32(12):1492-1499
  55. 55. Hnisz D et al. Activation of proto-oncogenes by disruption of chromosome neighborhoods. Science. 2016;351(6280):1454-1458
  56. 56. Parelho V et al. Cohesins functionally associate with CTCF on mammalian chromosome arms. Cell. 2008;132(3):422-433
  57. 57. Wendt KS et al. Cohesin mediates transcriptional insulation by CCCTC-binding factor. Nature. 2008;451(7180):796-801
  58. 58. Gruber S, Haering CH, Nasmyth K. Chromosomal cohesin forms a ring. Cell. 2003;112(6):765-777
  59. 59. Lara-Pezzi E et al. Evidence of a transcriptional co-activator function of cohesin STAG/SA/Scc3. The Journal of Biological Chemistry. 2004;279(8):6553-6559
  60. 60. Lengronne A et al. Cohesin relocation from sites of chromosomal loading to places of convergent transcription. Nature. 2004;430(6999):573-578
  61. 61. Busslinger GA et al. Cohesin is positioned in mammalian genomes by transcription, CTCF and Wapl. Nature. 2017;544(7651):503-507
  62. 62. Kueng S et al. Wapl controls the dynamic association of cohesin with chromatin. Cell. 2006;127(5):955-967
  63. 63. Zuin J et al. Cohesin and CTCF differentially affect chromatin architecture and gene expression in human cells. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(3):996-1001
  64. 64. Haarhuis JHI et al. The Cohesin release factor WAPL restricts chromatin loop extension. Cell. 2017;169(4):693-707 e14
  65. 65. Alipour E, Marko JF. Self-organization of domain structures by DNA-loop-extruding enzymes. Nucleic Acids Research. 2012;40(22):11202-11212
  66. 66. Fudenberg G et al. Formation of chromosomal domains by loop extrusion. Cell Reports. 2016;15(9):2038-2049
  67. 67. Phanstiel DH et al. Static and dynamic DNA loops form AP-1-bound activation hubs during macrophage development. Molecular Cell. 2017;67(6):1037-1048 e6
  68. 68. Javierre BM et al. Lineage-specific genome architecture links enhancers and non-coding disease variants to target gene promoters. Cell. 2016;167(5):1369-1384 e19
  69. 69. Hakim O et al. Glucocorticoid receptor activation of the Ciz1-Lcn2 locus by long range interactions. The Journal of Biological Chemistry. 2009;284(10):6048-6052
  70. 70. Hakim O et al. Diverse gene reprogramming events occur in the same spatial clusters of distal regulatory elements. Genome Research. 2011;21(5):697-706
  71. 71. Stavreva DA et al. Dynamics of chromatin accessibility and long-range interactions in response to glucocorticoid pulsing. Genome Research. 2015;25(6):845-857
  72. 72. Teferedegne B et al. Mechanism of action of a distal NF-kappaB-dependent enhancer. Molecular and Cellular Biology. 2006;26(15):5759-5770
  73. 73. Vernimmen D et al. Long-range chromosomal interactions regulate the timing of the transition between poised and active gene expression. The EMBO Journal. 2007;26(8):2041-2051
  74. 74. Andrey G et al. A switch between topological domains underlies HoxD genes collinearity in mouse limbs. Science. 2013;340(6137):1234167
  75. 75. Malovannaya A et al. Analysis of the human endogenous coregulator complexome. Cell. 2011;145(5):787-799
  76. 76. Yamamoto KR. Steroid receptor regulated transcription of specific genes and gene networks. Annual Review of Genetics. 1985;19:209-252
  77. 77. Luisi BF et al. Crystallographic analysis of the interaction of the glucocorticoid receptor with DNA. Nature. 1991;352(6335):497-505
  78. 78. Gronemeyer H, Bourguet W. Allosteric effects govern nuclear receptor action: DNA appears as a player. Science Signaling. 2009;2(73):pe34
  79. 79. Meijsing SH et al. DNA binding site sequence directs glucocorticoid receptor structure and activity. Science. 2009;324(5925):407-410
  80. 80. Watson LC et al. The glucocorticoid receptor dimer interface allosterically transmits sequence-specific DNA signals. Nature Structural & Molecular Biology. 2013;20(7):876-883
  81. 81. Presman DM et al. DNA binding triggers tetramerization of the glucocorticoid receptor in live cells. Proceedings of the National Academy of Sciences of the United States of America. 2016;113(29):8236-8241
  82. 82. Presman DM et al. Live cell imaging unveils multiple domain requirements for in vivo dimerization of the glucocorticoid receptor. PLoS Biology. 2014;12(3):e1001813
  83. 83. Starick SR et al. ChIP-exo signal associated with DNA-binding motifs provides insight into the genomic binding of the glucocorticoid receptor and cooperating transcription factors. Genome Research. 2015;25(6):825-835
  84. 84. Lim HW et al. Genomic redistribution of GR monomers and dimers mediates transcriptional response to exogenous glucocorticoid in vivo. Genome Research. 2015;25(6):836-844
  85. 85. Vinson C, Acharya A, Taparowsky EJ. Deciphering B-ZIP transcription factor interactions in vitro and in vivo. Biochimica et Biophysica Acta. 2006;1759(1-2):4-12
  86. 86. Cooper CD, Newman JA, Gileadi O. Recent advances in the structural molecular biology of Ets transcription factors: Interactions, interfaces and inhibition. Biochemical Society Transactions. 2014;42(1):130-138
  87. 87. Karin M. New twists in gene regulation by glucocorticoid receptor: Is DNA binding dispensable? Cell. 1998;93(4):487-490
  88. 88. Kassel O, Herrlich P. Crosstalk between the glucocorticoid receptor and other transcription factors: Molecular aspects. Molecular and Cellular Endocrinology. 2007;275(1-2):13-29
  89. 89. Baeuerle PA, Baltimore D. NF-kappa B: Ten years after. Cell. 1996;87(1):13-20
  90. 90. De Bosscher K, Vanden Berghe W, Haegeman G. The interplay between the glucocorticoid receptor and nuclear factor-kappaB or activator protein-1: Molecular mechanisms for gene repression. Endocrine Reviews. 2003;24(4):488-522
  91. 91. Hayden MS, Ghosh S. NF-kappaB, the first quarter-century: Remarkable progress and outstanding questions. Genes & Development. 2012;26(3):203-234
  92. 92. Natoli G. NF-kappaB and chromatin: Ten years on the path from basic mechanisms to candidate drugs. Immunological Reviews. 2012;246(1):183-192
  93. 93. Glass CK, Saijo K. Nuclear receptor transrepression pathways that regulate inflammation in macrophages and T cells. Nature Reviews. Immunology. 2010;10(5):365-376
  94. 94. Uhlenhaut NH et al. Insights into negative regulation by the glucocorticoid receptor from genome-wide profiling of inflammatory cistromes. Molecular Cell. 2013;49(1):158-171
  95. 95. Oh KS et al. Anti-inflammatory chromatinscape suggests alternative mechanisms of glucocorticoid receptor action. Immunity. 2017;47(2):298-309 e5
  96. 96. Kadiyala V et al. Cistrome-based cooperation between airway epithelial glucocorticoid receptor and NF-kappaB orchestrates anti-inflammatory effects. The Journal of Biological Chemistry. 2016;291(24):12673-12687
  97. 97. Weikum ER et al. Tethering not required: The glucocorticoid receptor binds directly to activator protein-1 recognition motifs to repress inflammatory genes. Nucleic Acids Research. 2017;45(14):8596-8608
  98. 98. Chinenov Y et al. Role of transcriptional coregulator GRIP1 in the anti-inflammatory actions of glucocorticoids. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(29):11776-11781
  99. 99. Logie C, Stewart AF. Ligand-regulated site-specific recombination. Proceedings of the National Academy of Sciences of the United States of America. 1995;92(13):5940-5944
  100. 100. Inayoshi Y et al. Repression of GR-mediated expression of the tryptophan oxygenase gene by the SWI/SNF complex during liver development. Journal of Biochemistry. 2005;138(4):457-465
  101. 101. Muratcioglu S et al. Structural modeling of GR interactions with the SWI/SNF chromatin remodeling complex and C/EBP. Biophysical Journal. 2015;109(6):1227-1239
  102. 102. Boldizsar F et al. Emerging pathways of non-genomic glucocorticoid (GC) signalling in T cells. Immunobiology. 2010;215(7):521-526
  103. 103. Lowenberg M et al. Novel insights into mechanisms of glucocorticoid action and the development of new glucocorticoid receptor ligands. Steroids. 2008;73(9-10):1025-1029
  104. 104. Buttgereit F, Scheffold A. Rapid glucocorticoid effects on immune cells. Steroids. 2002;67(6):529-534
  105. 105. Beck IM et al. Crosstalk in inflammation: The interplay of glucocorticoid receptor-based mechanisms and kinases and phosphatases. Endocrine Reviews. 2009;30(7):830-882
  106. 106. Sionov RV et al. Role of mitochondrial glucocorticoid receptor in glucocorticoid-induced apoptosis. The Journal of Experimental Medicine. 2006;203(1):189-201
  107. 107. Stellato C. Post-transcriptional and nongenomic effects of glucocorticoids. Proceedings of the American Thoracic Society. 2004;1(3):255-263
  108. 108. Cruz-Topete D, Cidlowski JA. One hormone, two actions: Anti- and pro-inflammatory effects of glucocorticoids. Neuroimmunomodulation. 2015;22(1-2):20-32
  109. 109. Battich N, Stoeger T, Pelkmans L. Control of transcript variability in single mammalian cells. Cell. 2015;163(7):1596-1610
  110. 110. Kino T et al. Noncoding RNA gas5 is a growth arrest- and starvation-associated repressor of the glucocorticoid receptor. Science Signaling. 2010;3(107):ra8
  111. 111. Tani H, Torimura M, Akimitsu N. The RNA degradation pathway regulates the function of GAS5 a non-coding RNA in mammalian cells. PLoS One. 2013;8(1):e55684
  112. 112. Strahle U et al. Glucocorticoid- and progesterone-specific effects are determined by differential expression of the respective hormone receptors. Nature. 1989;339(6226):629-632
  113. 113. Mayama T, Marr AK, Kino T. Differential expression of glucocorticoid receptor noncoding RNA repressor Gas5 in autoimmune and inflammatory diseases. Hormone and Metabolic Research. 2016;48(8):550-557
  114. 114. Lucafo M et al. Role of the long non-coding RNA growth arrest-specific 5 in glucocorticoid response in children with inflammatory bowel disease. Basic & Clinical Pharmacology & Toxicology. 2017
  115. 115. Page MF et al. SMG-2 is a phosphorylated protein required for mRNA surveillance in Caenorhabditis elegans and related to Upf1p of yeast. Molecular and Cellular Biology. 1999;19(9):5943-5951
  116. 116. Bhattacharya A et al. Characterization of the biochemical properties of the human Upf1 gene product that is involved in nonsense-mediated mRNA decay. RNA. 2000;6(9):1226-1235
  117. 117. Shyu AB, Wilkinson MF. The double lives of shuttling mRNA binding proteins. Cell. 2000;102(2):135-138
  118. 118. Anderson P, Kedersha N. RNA granules: Post-transcriptional and epigenetic modulators of gene expression. Nature Reviews. Molecular Cell Biology. 2009;10(6):430-436
  119. 119. Chen CY, Shyu AB. AU-rich elements: Characterization and importance in mRNA degradation. Trends in Biochemical Sciences. 1995;20(11):465-470
  120. 120. Chen CY et al. Nucleolin and YB-1 are required for JNK-mediated interleukin-2 mRNA stabilization during T-cell activation. Genes & Development. 2000;14(10):1236-1248
  121. 121. Anderson P. Post-transcriptional control of cytokine production. Nature Immunology. 2008;9(4):353-359
  122. 122. Ishmael FT et al. Role of the RNA-binding protein tristetraprolin in glucocorticoid-mediated gene regulation. Journal of Immunology. 2008;180(12):8342-8353
  123. 123. Kratochvill F et al. Tristetraprolin limits inflammatory cytokine production in tumor-associated macrophages in an mRNA decay-independent manner. Cancer Research. 2015;75(15):3054-3064
  124. 124. Poon M, Liu B, Taubman MB. Identification of a novel dexamethasone-sensitive RNA-destabilizing region on rat monocyte chemoattractant protein 1 mRNA. Molecular and Cellular Biology. 1999;19(10):6471-6478
  125. 125. Dhawan L et al. A novel role for the glucocorticoid receptor in the regulation of monocyte chemoattractant protein-1 mRNA stability. The Journal of Biological Chemistry. 2007;282(14):10146-10152
  126. 126. Ishmael FT et al. The human glucocorticoid receptor as an RNA-binding protein: Global analysis of glucocorticoid receptor-associated transcripts and identification of a target RNA motif. Journal of Immunology. 2011;186(2):1189-1198
  127. 127. Cho H et al. Glucocorticoid receptor interacts with PNRC2 in a ligand-dependent manner to recruit UPF1 for rapid mRNA degradation. Proceedings of the National Academy of Sciences of the United States of America. 2015;112(13):E1540-E1549
  128. 128. Park OH, Do E, Kim YK. A new function of glucocorticoid receptor: Regulation of mRNA stability. BMB Reports. 2015;48(7):367-368
  129. 129. Park OH et al. Identification and molecular characterization of cellular factors required for glucocorticoid receptor-mediated mRNA decay. Genes & Development. 2016;30(18):2093-2105
  130. 130. Rabani M et al. High-resolution sequencing and modeling identifies distinct dynamic RNA regulatory strategies. Cell. 2014;159(7):1698-1710
  131. 131. Wang H et al. The effects of microRNAs on glucocorticoid responsiveness. Journal of Cancer Research and Clinical Oncology. 2017;143(6):1005-1011

Written By

Cheng Wang, Roel Oldenkamp, Ronald J.W. Oellers and Colin Logie

Submitted: 01 June 2017 Reviewed: 27 October 2017 Published: 20 December 2017